首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Ground-state spin-Hamiltonian parameters, magnetic data, and electronic Raman spectra of hexacoordinate vanadium(III) complexes are calculated with consideration to the ((3)A (3)E) e vibronic interaction and compared to experimental data. It is shown that the zero-field-splitting of the (3)A(g) (S(6)) ground term may be reduced significantly by the dynamical Jahn-Teller effect, particularly when the pi-anisotropy of the metal-ligand bonding interaction is significant, and the energy of the Jahn-Teller active vibration is comparable to the diagonal axial field. The dynamical Jahn-Teller effect may also give rise to a significant enhancement in the Raman intensity of overtones and higher harmonics of Jahn-Teller active vibrations, when the energies of these transitions fall in the proximity of intra-(3)T(1g) (O(h)) electronic Raman transitions. A simple method of conducting vibronic coupling calculations is described, employing ligand field matrices generated by angular overlap model calculations, which may in principle be applied to any transition metal complex.  相似文献   

2.
Molecular magnetism in a series of cyano‐bridged first and second transition metal complexes has been investigated using density functional theory (DFT) combined with the broken‐symmetry (BS) approach. Several exchange‐correlation (XC) functionals in the ADF package were used to investigate complexes I [?(Me3tacn)2(cyclam)NiMo2(CN)6]2+, II [?(Me3tacn)2(cyclam)Ni‐Cr2(CN)6]2+, III [(Me3tacn)6MnMo6(CN)18]2+, and IV [(Me3tacn)6MnCr6(CN)18]2+ (Me3tacn = N,N′,N?‐trimethyl‐1,4,7‐triazacyclononane). For models A (the molded structure of complex I) and B (the modeled structure of complex II), all the XCs given qualitatively reasonable results and predict ferromagnetic coupling character between M (M = MoIII for A or CrIII for B) and NiII in coincidence with the experimental results (see Tables I and II ). The calculated using Operdew, OPBE, O3LYP, and B3LYP functionals and experimental J values show that substituting CrIII with MoIII will enhance the ferromagnetic exchange coupling interactions. But VWN, PW91, PBE, VSXC, and tau‐HCTH functionals have no way to differentiate the relative strength of the intramolecular magnetic exchange coupling interactions of A and B correctly. For models C (the modeled structure of complex III) and D (the modeled structure of complex IV), all the XCs in ADF and B3LYP in Gaussian 03 with several basis sets show that substituting CrIII with MoIII will enhance the antiferromagnetic exchange coupling interactions. From the above calculations, the substitution of CrIII by MoIII will enhance the magnetic coupling interactions, whether the magnetic coupling interactions are ferro‐ or antiferromagnetic. Moreover, Kahn's model was applied to investigate the above facts. © 2006 Wiley Periodicals, Inc. Int J Quantum Chem, 2006  相似文献   

3.
4.
Photoluminescence has been observed from a series of Zn(II) and Cd(II) complexes containing both N-heterocyclic and aromatic thiol ligands. The emission consists of overlapping bands, a structured phosphorescence that arises from the lowest 3ππ* state of the heterocycle and a broad band luminescence that maximizes at still lower energies and decays on the order of 10–50 μs. This new band is assigned to an interligand charge-transfer excited configuration in which electronic charge has been transferred from the coordinated thiol ligand(s) to the N-heterocycle via the metal ion. A superexchange mechanism is proposed to account for the disposition of singlet and triplet states.  相似文献   

5.
The preparation and properties of titanium(III) cationic complexes with hexamethylbenzene are described.Hexamethylbenzene was found to form monometallic cationic complexes [(HMBTiC12)+] as well as trimetallic cationic clusters [(HMB3Ti3C16)+]. Composition of the latter however remains uncertain.The ability of HMB to form complexes with titanium(III) appears to be unique in the series of polymethylbenzenes.  相似文献   

6.
7.
Iron(III) complexes with the general formula of [Fe (R-Himap)2]X and [Fe(R-Himat)2]X (Himap = shiff-base prepared from the condensation of 4-formylimidazole and 2-aminophenol, Himat = shiff base prepared from the condensation of 4-formylimidazole and 2-aminobenzenethiol, and R = H, Me, Ph; X = ClO4, NO3, BPh4) have been synthesized. The complexes have an N4O2 donor set or an N4S2 donor set. These complexes have 5 and 5 member rings around an iron(III) atom per one chelate ring, that is, "5-5 member rings". The crystal structure, Mössbauer spectra, magnetic properties and absorption spectra of the complexes were examined. In addition, [Fe (Himsa)2]ClO4 having "5-6 member rings" of an N4O2 donor set and [Fe (Ph-Himap)(Ph-imap)] obtained by the deprotonation of [Fe (Ph-Himap)2]ClO4 have been also synthesized, and the Mössbauer spectra and magnetic properties of those complexes were examined. The X-ray structure of a single crystal of [Fe (Himap)2]BPh4 was determined: C44H36N6BO2Fe, triclinic, space group P(# 2), a = 12.452(2) Å, b = 12.748(2) Å, c = 11.996(2) Å, = 103.97(1)°, = 90.78(1)°, = 84.70(1)° and Z = 2. The moiety of an iron atom of [Fe(Himap)2]BPh4 was a pseudo octahedron with an FeN4O2 geometry. [Fe(R-Himap)2]X was in high-spin state (about 5.9 B.M. at 80 K in solid state), and [Fe(R-Himat)2]X was in low-spin state (about 2.0 B.M. at 293 K). The complex [Fe(Himsa)2]ClO4 with "5-6 member rings" and the deprotonated complex [Fe(Ph-Himap)(Ph-imap)] were in the high-spin state (6.0 B.M. at 80 K). It is concluded that the ligand field strength of an N4S2 donor set is stronger than that of an N4O2 donor set.  相似文献   

8.
Density functional theory is carried out to study hexaphyrin and its bis-metal and mixed bis-metal (M = Cu3+, Ag3+, and Au3+) complexes. The electronic structures and bonding situations of them are studied by using natural bond orbital approach and the topological analysis of the electron localization function. Electronic spectra are investigated by using time-dependent density functional theory. The introduction of group 11 transition metals leads to red shifts in the spectra of these metal complexes with respect to that of hexaphyrin. Moreover, it is noteworthy that the spectra of copper contained complexes are mainly derived from combination of ligand-to-metal charge transfer and ligand-to-ligand charge transfer transitions. In addition, the relativistic time-dependent density functional theory with spin-orbit coupling calculations indicate that the effects of spin-orbit coupling on the excitation energies are so small that it is safe enough to neglect spin-orbit coupling for these systems.  相似文献   

9.
Spectroscopic and crystallographic data are presented for salts containing the [V(OH(2))(6)](3+) cation, providing a rigorous test of the ability of the angular overlap model (AOM) to inter-relate the electronic and molecular structure of integer-spin complexes. High-field multifrequency EPR provides a very precise definition of the ground-state spin-Hamiltonian parameters, while single-crystal absorption measurements enable the energies of excited ligand-field states to be identified. The EPR study of vanadium(III) as an impurity in guanidinium gallium sulfate is particularly instructive, with fine-structure observed attributable to crystallographically distinct [V(OH(2))(6)](3+) cations, hyperfine coupling, and ferroelectric domains. The electronic structure of the complex depends strongly on the mode of coordination of the water molecules to the vanadium(III) cation, as revealed by single-crystal neutron and X-ray diffraction measurements, and is also sensitive to the isotopic abundance. It is shown that the AOM gives a very good account of the change in the electronic structure, as a function of geometric coordinates of the [V(OH(2))(6)](3+) cation. However, the ligand-field analysis is inconsistent with the profiles of electronic transitions between ligand-field terms.  相似文献   

10.
Single-crystal Raman spectra are presented for the salt [C(NH2)3][V(OH2)6](SO4)2, displaying electronic transitions between the trigonal components of the vanadium(III) 3T1g(Oh) ground term. The 3A-->3E(C3) electronic Raman band is centered at approximately 2720 cm-1, and exhibits extensive structure, revealing the energies of the spinor components of the 3E(C3) term for the two crystallographically distinct [V(OH2)6]3+ cations. The data are interpreted in conjunction with parameters previously reported from an electron paramagnetic resonance study of the salt. A satisfactory reproduction of the electronic Raman profile and ground-state spin-Hamiltonian parameters is achieved by employing a (3A plus sign in circle3E)multiply sign in circle e vibronic coupling model, in which the spin-orbit splitting of the 3E(C3) is quenched significantly by the Ham effect, and the intensity of harmonics of the Jahn-Teller active vibration enhanced by their proximity to the electronic Raman bands. The model gives an excellent account of the intensities of the electronic Raman bands, which are shown to depend profoundly on both temperature and the selected component of the polarizability tensor. The electronic Raman profile changes notably upon deuteriation, a result that exposes deficiencies in the single-mode coupling model.  相似文献   

11.
A variety of physical methods has been used to probe the non-Kramers, S = 1, V(III) ion in two types of pseudooctahedral complexes: V(acac)(3), where acac = anion of 2,4-pentanedione, and VX(3)(thf)(3), where thf = tetrahydrofuran and X = Cl and Br. These methods include tunable frequency and high-field electron paramagnetic resonance (HFEPR) spectroscopy (using frequencies of approximately 95-700 GHz and fields up to 25 T) in conjunction with electronic absorption, magnetic circular dichroism (MCD), and variable-temperature variable-field MCD (VTVH-MCD) spectroscopies. Variable-temperature magnetic susceptibility and field-dependent magnetization measurements were also performed. All measurements were conducted on complexes in the solid state (powder or mull samples). The field versus sub-THz wave quantum energy dependence of observed HFEPR resonances yielded the following spin Hamiltonian parameters for V(acac)(3): D = +7.470(1) cm(-1); E = +1.916(1) cm(-1); g(x) = 1.833(4); g(y) = 1.72(2); g(z) = 2.03(2). For VCl(3)(thf)(3), HFEPR detected a single zero-field transition at 15.8 cm(-1) (474 GHz), which was insufficient to determine the complete set of spin Hamiltonian parameters. For VBr(3)(thf)(3), however, a particularly rich data set was obtained using tunable-frequency HFEPR, and analysis of this data set gave the folowing: D = -16.162(6) cm(-1); E = -3.694(4) cm(-1); g(x) = 1.86(1); g(y) = 1.90(1); g(z) = 1.710(4). Analysis of the VTVH-MCD data gave spin Hamiltonian parameters in good agreement with those determined by HFEPR for both V(acac)(3) and VBr(3)(thf)(3) and in rough agreement with the estimate for VCl(3)(thf)(3) (D approximately 10 cm(-1), |E/D| approximately 0.18), together with the finding that the value of D is negative for both thf complexes. The electronic structures of these V(III) complexes are discussed in terms of their molecular structures and the electronic transitions observed by electronic absorption and MCD spectroscopies.  相似文献   

12.
The [Os(III)(CN)6]3- anion is prepared by chemical oxidation in aqueous solution and isolated as yellow prisms of [Ph4P]3[Os(III)(CN)6].6H2O (1). This species crystallizes in the triclinic space group P with cell parameters a = 13.7609(11) A, b = 16.2275(13) A, c = 17.0895(14) A, alpha = 91.4040(10) degrees , beta = 109.3600(10) degrees , gamma = 102.3970(10) degrees , V = 3497.4(5) A(3), and Z = 2. The slightly distorted octahedral moiety displays Os-C and C-N bond lengths that average 2.058 and 1.146 A, respectively. Spin-orbit-coupling splitting of the ground-state term dominates the NIR region of the electronic spectrum and the magnetic behavior of 1. The experimental information points to higher spin delocalization over the coordinated cyanides than in [Fe(III)(CN)6]3-.  相似文献   

13.
Two ruthenium(III) complexes containing ethylenediaminetetraacetate(edta), viz. [{Ru(Hedta)}2L]·xH2O L = 4,4′-bipyridine(bpy) (1) and 4,4′-azopyridine(Azpy) (2), have been synthesized by the reaction between K[Ru(Hedta)Cl]·1.5H2O and the corresponding N-heterocycles. Complex 1 was determined by single-crystal X-ray diffraction. The products were characterized by IR, UV–vis, cyclic voltammetry, and magnetic techniques. Their DNA-binding activities were investigated using electronic absorption spectroscopic methods and ?uorescence quenching; the experimental results show that these two ruthenium complexes may bind to CT-DNA through intercalation modes.  相似文献   

14.
The reaction of 2,6-dimethoxypyridine-3-carboxylic acid (DMPH) with different precursors [Ti(η5-C5H5)2Cl2], [Ti(η5-C5H4Me)2Cl2], [Ti(η5-C5H4SiMe3)(η5-C5H5)Cl2], [Ti(η5-C5Me5)Cl3], SnMe3Cl and GatBu3 yielded the complexes [Ti(η5-C5H5)2(DMP-κO)2] (1), [Ti(η5-C5H4Me)2(DMP-κO)2] (2), [Ti(η5-C5H4SiMe3)(η5-C5H5)(DMP-κO)2] (3), [Ti(η5-C5Me5)(DMP-κ2O,O′)3] (4), [SnMe3(μ-DMP-κOO′)] (5), and [GatBu2(μ-DMP-κOO′)]2 (6). 1-6 have been characterized by spectroscopic methods and the molecular structure of the complexes 1, 2, 3, 5 and 6 have been determined by X-ray diffraction studies. The cytotoxic activity of 1-6 was tested against the tumour cell lines human adenocarcinoma HeLa, human myelogenous leukaemia K562, human malignant melanoma Fem-x and human breast carcinoma MDA-MB-361. The results of this study show a higher cytotoxicity of the tin(IV) and gallium(III) derivatives in comparison to their titanium(IV) counterparts. Furthermore, the different titanium compounds showed differences in their cytotoxicities with a higher activity of complex 4 (mono-(cyclopentadienyl) derivative) compared to that of 1-3 (bis-(cyclopentadienyl) complexes). A qualitative UV-vis study of the interactions of these complexes with DNA has also been carried out.  相似文献   

15.
Chemically reasonable models of PR3 (R = Me, Et, iPr, and tBu) were constructed to apply the post Hartree-Fock method to large transition metal complexes. In this model, R is replaced by the H atom including the frontier orbital consistent quantum capping potential (FOC-QCP) which reproduces the frontier orbital energy of PR3. The steric effect is incorporated by the new procedure named steric repulsion correction (SRC). To examine the performance of this FOC-QCP method with the SRC, the activation barriers and reaction energies of the reductive elimination reactions of C2H6 and H2 from M(R1)2(PR2(3))2 (M = Ni, Pd, or Pt; R1 = Me for R2 = Me, Et, or iPr, or R1 = H for R2 = tBu) were evaluated with the DFT[B3PW91], MP4(SDQ), and CCSD(T) methods. The FOC-QCP method reproduced well the DFT[B3PW91]- and MP4(SDQ)-calculated energy changes of the real complexes with PMe3. For more bulky phosphine, the SRC is important to present correct energy change, in which the MP2 method presents reliable steric repulsion correction like the CCSD(T) method because the systems calculated in the SRC do not include a transition metal element. The monomerization energy of [RhCl(PiPr3)2]2 and the coordination energies of CO, H2, N2, and C2H4 with [RhCl(PiPr3)2]2 were theoretically calculated by the CCSD(T) method combined with the FOC-QCP and the SRC. The CCSD(T)-calculated energies agree well with the experimental ones, indicating the excellent performance of the combination of the FOC-QCP with the SRC. On the other hand, the DFT[B3PW91]-calculated energies of the real complexes considerably deviate from the experimental ones.  相似文献   

16.
17.
The kinetics of oxidation of malic acid by peroxomonosulphate (PMS) in the presence of Cu(II) (2.50 × 10?4–5.00 × 10?3 M), Co(II) (2.00 × 10?6–1.00 × 10?5 M) and Ni(II) (5.00 × 10?4–6.00 × 10?3 M) were studied in the pH range 4.05–5.89. The oxidation of Ni(II) malate follows simple first-order kinetics with respect to both [PMS] and [Ni(II)], while the oxidations of Cu(II) malate and Co(II) malate show autocatalysis. There is an appreciable induction period in the Cu(II) malate oxidation, while Co(II) malate oxidation follows a simple curve. The initial oxidation product for all three systems was identified as malonic semialdehyde. Alcohol quenching studies suggest that, even in the Co(II) malate-PMS system, no radical intermediates such as $ {\text{SO}}_{4}^{ - .} $ or $ {\text{OH}}{}^{.} $ are detected. The malonic semialdehyde intermediate may react with M(II) malates to give a hemiacetal, which may be more reactive.  相似文献   

18.
High-pressure magnetic susceptibility measurements have been carried out on Fe(dipy)2(NCS)2 and Fe(phen)2(NCS)2 in the pressure range 1–10 kbar and tempeature range 80–300 K in order to investigate the factors responsible for the spin-state transitions. The transitions change from first order to second or higher order upon application of pressure. The temperature variation of the susceptibility at different pressures has been analysed quantitatively within the framework of available models. It is shown that the relative magnitudes of the ΔG0 of high-spin and low-spin conversion and the ferromagnetic interaction between high-spin complexes determines the nature of the transition.  相似文献   

19.
Summary The platinum(II) carboxylates,trans-Pt(O2CR)2(py)2 and Pt(O2CR)2bpy (R=C6F5,p-HC6F4,m-HC6F4, oro-HC6F4; bpy=2,2-bipyridyl), have been prepared by reactions oftrans-Pt(OH)2(py)2 or Pt(OH)2bpy with the appropriate polyfluorobenzoic acids, whilst [Pt(py)4](O2CC6F5)2 has been obtained from reaction oftrans-PtCl2(py)2 with thallous pentafluorobenzoate in pyridine at room temperature. In boiling pyridine, the platinum(II) polyfluorobenzoates undergo either decarboxylation givingtrans-PtR2(py)2 and PtR2bpy (R= C6F5,p-HC6F4, orm-HC6F4) complexes or substitution, giving [Pt(py)4](O2CC6F4H-o)2 and [Ptbpy(py)2](O2CC6F4H-o)2. Reactions oftrans-PtX2(py)2 and PtX2bpy (X=Cl or Br) with appropriate thallous polyfluorobenzoates in boiling pyridine have yielded the complexestrans-PtR2(py)2, PtR2bpy, PtCl(R)bpy (R=C6F5,p-HC6F4, orm-HC6F4 in each case),trans-PtCl(R)(py)2 (R = C6F5 orm-HC6F4),trans-PtBr(C6F5)(py)2, and PtBr(C6F5)bpy. The complexestrans-PtR2(py)2 (R=C6F5 orp-HC6F4) have also been prepared from potassium tetrachloroplatinate(II) and the appropriate thallous polyfluorobenzoate in boiling py, andtrans-Pt(C6F5)2(py)2 has been similarly obtained fromcis-PtCl2(py)2 and C6F5CO2Tl. Significant decarboxylation was not observed on reaction oftrans-PtCl2(py)2 or PtCl2bpy with thallous 2,3,4,5-tetrafluorobenzoate.Part II, ref. 4;Preliminary communication, ref. 3;  相似文献   

20.
《Polyhedron》1987,6(5):1009-1015
Reactions of 2-mercapto-3-phenyl-4-quinazolinone (LH) with RuCl3·xH2O and RhCl3·xH2O afforded the compounds [RuL2Cl(H2O)]H2O, [RuL2Cl·DMFI and RhL(LH)Cl2·2H2O. Reactions of LH with RuCl3·xH2O in the presence of N-heterocyclic bases led to the formation of complexes of type [RuL2ClB]·H2O (B = pyridine, 3-picoline or imidazole) and [RuLCl2(o-phen)] H2O (o-phen = 1, 10-phenanthroline). These complexes were characterized on the basis of analytical, conductivity, magnetic, IR and electronic spectral and ESR studies. Tentative structures for the complexes are proposed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号